32
X-ray Scintillation in Lead Halide Perovskite Crystals M. D. Birowosuto, 1,2,* D. Cortecchia, 3,4 W. Drozdowski, 5 K. Brylew, 5 W. Lachmanski, 5 A. Bruno, 4 C. Soci 2,4,6,* 1 CINTRA UMI CNRS/NTU/THALES 3288, Research Techno Plaza, 50 Nanyang Drive, Border X Block, Level 6, Singapore 637553 2 Center of Disruptive Photonic Technologies, TPI, SPMS, Nanyang Technological University, 21 Nanyang Link, Singapore 637371 3 Interdisciplinary Graduate School, Nanyang Technological University, Singapore 639798 4 Energy Research Institute @ NTU (ERI@N), Research Techno Plaza, Nanyang Technological University, 50 Nanyang Drive, Singapore 637553 5 Institute of Physics, Faculty of Physics, Astronomy, and Informatics, Nicolaus Copernicus University, Grudziadzka 5, 87-100 Torun, Poland 6 School of Physical and Mathematical Sciences, Division of Physics and Applied Physics, Nanyang Technological University, 21 Nanyang Link, Singapore 637371 *Corresponding authors: [email protected]; [email protected] Keywords: X-ray scintillators, organometallic perovskites, single crystals, thermoluminescence

Lead Halide Perovskite Crystals for X-ray Scintillators · applications, from crystallography [8] to space exploration [9]. Modern X-ray detectors rely on two main mechanisms of energy

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

  • X-ray Scintillation in Lead Halide Perovskite Crystals

    M. D. Birowosuto,1,2,* D. Cortecchia,3,4 W. Drozdowski,5 K. Brylew,5 W. Lachmanski,5

    A. Bruno,4 C. Soci2,4,6,*

    1CINTRA UMI CNRS/NTU/THALES 3288, Research Techno Plaza, 50 Nanyang Drive,

    Border X Block, Level 6, Singapore 637553

    2Center of Disruptive Photonic Technologies, TPI, SPMS, Nanyang Technological

    University, 21 Nanyang Link, Singapore 637371

    3Interdisciplinary Graduate School, Nanyang Technological University, Singapore 639798

    4 Energy Research Institute @ NTU (ERI@N), Research Techno Plaza, Nanyang

    Technological University, 50 Nanyang Drive, Singapore 637553

    5Institute of Physics, Faculty of Physics, Astronomy, and Informatics, Nicolaus Copernicus

    University, Grudziadzka 5, 87-100 Torun, Poland

    6 School of Physical and Mathematical Sciences, Division of Physics and Applied Physics,

    Nanyang Technological University, 21 Nanyang Link, Singapore 637371

    *Corresponding authors: [email protected]; [email protected]

    Keywords: X-ray scintillators, organometallic perovskites, single crystals, thermoluminescence

    mailto:[email protected]:[email protected]

  • 1

    Abstract

    Current technologies for X-ray detection rely on scintillation from expensive inorganic crystals

    grown at high-temperature, which so far has hindered the development of large-area scintillator

    arrays. Thanks to the presence of heavy atoms, solution-grown hybrid lead halide perovskite

    single crystals exhibit short X-ray absorption length and excellent detection efficiency. Here we

    compare X-ray scintillator characteristics of three-dimensional (3D) MAPbI3 and MAPbBr3 and

    two-dimensional (2D) (EDBE)PbCl4 hybrid perovskite crystals. X-ray excited

    thermoluminescence measurements indicate the absence of deep traps and a very small density

    of shallow trap states, which lessens after-glow effects. All perovskite single crystals exhibit

    high X-ray excited luminescence yields of >120,000 photons/MeV at low temperature. Although

    thermal quenching is significant at room temperature, the large exciton binding energy of 2D

    (EDBE)PbCl4 significantly reduces thermal effects compared to 3D perovskites, and moderate

    light yield of 9,000 photons/MeV can be achieved even at room temperature. This highlights the

    potential of 2D metal halide perovskites for large-area and low-cost scintillator devices for

    medical, security and scientific applications.

  • 2

    Introduction

    The investigation of X-ray detectors started with the discovery of X-rays by Wilhelm Röntgen,

    who noticed the glow from a barium platino-cyanide screen placed besides a vacuum tube [1,2].

    Since this discovery, more than one hundred years ago, the development of efficient [3-5] and

    large-area [5-7] X-ray detectors has been a topic of continuous interest, targeting a wide range of

    applications, from crystallography [8] to space exploration [9].

    Modern X-ray detectors rely on two main mechanisms of energy conversion. The first is

    photon-to-current conversion, in which a semiconducting material directly converts the incoming

    radiation into electrical current [4-6]; the second is X-ray to UV-visible photon down-

    conversion, in which a scintillator material is coupled to a sensitive photodetector operating at

    lower photon energies [2]. Both methods are equally compelling for practical implementations,

    although their viability will ultimately depend on the development of new materials to overcome

    some of the current limitations, such as high cost, small area, and low conversion efficiency of

    the X-ray absorbers. Recent demonstrations of the use of hybrid metal-halide perovskites for X-

    and -ray detection has spurred great interest in this class of materials [7,10,11,12]. Besides their

    good detection efficiency, solution processing holds great promise for facile integration and

    development of industrial and biomedical applications.

    Methylammonium lead trihalide perovskites (MAPbX3 where MA=CH3NH3 and X= I, Br, or

    Cl) have demonstrated excellent performance in optoelectronic devices like field effect

    transistors [13], highly sensitive photodetectors for visible region [14], and light emitting devices

    [15,16]. Moreover, compositional tuning was used to realize tunable-wavelength lasers [17]. As

    X-ray detectors, MAPbX3 yield notably large X-ray absorption cross section due to large atomic

    numbers of the heavy Pb and I, Br, Cl atoms [10,11]. Thin-film MAPbX3 p-i-n photodiode and

  • 3

    lateral photoconductor devices have shown good efficiency for X-ray photon-to-current

    conversion [10,11]. However, thin-film X-ray detectors have typically low responsivity at high

    (keV) photon energies, where the absorption length (~mm) is much larger than the film thickness

    (~μm); even if thickness is increased to improve detection probability, direct photon-to-current

    conversion is ultimately hampered by the limited carrier-diffusion length (~1 μm in perovskites)

    [10]. Efficient X-ray photon-to-current conversion has been shown recently in single-crystal

    (thick) perovskite MAPbBr3, but sensitivity is still limited to energies up to 50 keV [11]. Also,

    standard -photon counting for energies up to 662 keV has been demonstrated in MAPbI3 [12].

    As opposed to direct photon-to-current conversion detectors, X-ray scintillators do not suffer

    from limited carrier diffusion length of the absorbing material [18,19]. Thin films of

    phenethylammonium lead bromide, PhE-PbBr4, with sub-nanosecond scintillation decay time

    have been previously tested in X-ray [20] and proton [21] scintillators, but yielded only 5-6%

    detection efficiency of 60 keV X-rays, limited by the film thickness (200 μm) [21]. By

    combining the good high-energy response with large absorption cross section deriving from

    large thickness and high mass-density, single crystal perovskite scintillators are therefore

    expected to improve detection efficiency of keV X- or γ-rays.

    In this paper, we present a thorough comparative study of the scintillation properties of three-

    dimensional (3D) and two-dimensional (2D) low-bandgap perovskite single crystals. We have

    synthesized mm-scale 3D perovskite crystals MAPbI3 and MAPbBr3, and 2D perovskite crystal

    (EDBE)PbCl4 (EDBE=2,2’-(ethylenedioxy)bis(ethylammonium)), comprising of alternating

    organic and inorganic layers which form a multi-quantum-well-like structure. The excellent

    quality of these crystals is indicated by structural analysis and by the very small density of

    shallow traps (n0~105-107 cm-3, E~10-90 meV) determined by X-ray excited

  • 4

    thermoluminescence, which reduces after-glow effects. Thanks to their lower bandgap compared

    to traditional scintillator crystals [6], perovskite crystals produce extremely high light yields of

    >120,000 photons/MeV (as estimated from X-ray-excited luminescence) at low temperature. In

    3D perovskites, the light yield is greatly reduced at room temperature (

  • 5

    MAPbBr3, and (EDBE)PbCl4 crystals appear lustrous black, orange, and white, respectively. The

    corresponding glows under ultraviolet lamp excitation are green and white for MAPbBr3 and

    (EDBE)PbCl4 crystals, while the glow of MAPbI3 could not be observed since its emission lies

    in the near infrared. Crystal colors and glows agree well with the absorption and

    photoluminescence properties of the corresponding thin films, which show optical energy gaps

    of Eg=1.51, 2.18, and 3.45 eV for MAPbI3 [22], MAPbBr3 and (EDBE)PbCl4 [24], respectively

    (Supplementary Fig. S4).

    Perovskite crystals offer multiple advantages for X-ray scintillation, specifically: i. Since the

    light yield of X-ray scintillation is inversely proportional to the optical bandgap Eg [2,18], low-

    bandgap perovskites of MAPbI3, MAPbBr3 and (EDBE)PbCl4 are expected to yield up to about

    270,000, 190,000, and 120,000 photons/MeV, respectively. Those light yields are much higher

    than state-of-art cerium (Ce3+) doped lanthanum tribromides LaBr3 (Eg=5.90 eV) [25,26] and

    Ce3 + doped lutetium iodides LuI3 (Eg=4.15 eV) [27,28] scintillators, with light yields of 68,000

    and 100,000 photons/MeV, respectively. ii. Since X-ray absorption length scales with the

    effective atomic number Zeff and mass density ρ [2], MAPbI3, MAPbBr3, and (EDBE)PbCl4

    (Zeff=66.83, 67.13, and 67.52, ρ=3.947, 3.582, and 2.191 gr⁄cm3, respectively) should reach X-

    ray absorption lengths up to 1 cm at 1 MeV, similar to Ce3+-doped LaBr3 and LuI3 scintillators

    (see Supplementary Fig. S5). iii. The unusually large Stokes shift of two-dimensional

    (EDBE)PbCl4 could be particularly beneficial to the scintillation yield [29], which is

    substantially reduced by self-absorption of the luminescence [30]. iv. The extremely fast

    photoluminescence decay of MAPbI3, MAPbBr3, and (EDBE)PbCl4 (fast components of 4.3,

    0.8-5.2, and 7.9 ns, respectively) may provide faster scintillation than 15 ns of Ce3+-doped LaBr3

    [25,30] and 33 ns of Ce3+-doped LuI3 [28] (Supplementary Fig. S6). Nanosecond scintillation

  • 6

    decay times were indeed demonstrated in PhE-PbBr4 using X-ray and -ray pulses, consistent

    with time-resolved photoluminescence [20,31]. v. Finally, long emission wavelengths in the

    range of 400 to 700 nm allow optimal detection of scintillation using highly sensitive avalanche

    photodiodes (APDs), which can reach quantum efficiencies up to 90-100% in comparison with

    photomultipliers (PMTs) with only 40-50% efficiency [28].

    Figure 1. Crystal structure and appearance. Top row: crystal structure

    representation of MAPbX3 (X=I, Br) three-dimensional perovskites (left), and

    (EDBE)PbCl4 two-dimensional perovskite (right); Middle row: photographs of the large

    single crystals of hybrid lead halide perovskites; Bottom row: glow of the crystals under

    ultraviolet lamp excitation. Scale bars: 5 mm.

  • 7

    Figure 2. Emission spectra under X-ray and optical excitation. X-ray excited

    luminescence (light color area) and photoluminescence (dark color area) spectra of a)

    MAPbI3, b) MAPbBr3, and c) (EDBE)PbCl4 recorded at room temperature with

    excitation wavelengths for photoluminescence of 425, 500, and 330 nm, respectively.

    Photoluminescence and X-ray excited luminescence spectra were normalized to their

    maxima, and normalized X-ray excited luminescence spectra were divided by a factor of

    two for clarity.

    The X-ray excited luminescence and photoluminescence spectra of MAPbI3, MAPbBr3, and

    (EDBE)PbCl4 crystals recorded at room temperature are shown in Fig. 2 (see experimental

    details in Materials and methods section). Both X-ray excited luminescence and

    photoluminescence spectra of MAPbI3 have a single broadband peak centered at 750 nm with

  • 8

    FWHM of ~80 nm (Fig. 2a). For MAPbBr3, both X-ray excited luminescence and

    photoluminescence spectra exhibit double peaks centered around 560 and 550 nm, respectively.

    MAPbBr3 has the narrowest emission band with full width of half-maximum (FWHM) of ~40

    nm (Fig. 2b). On the other hand, (EDBE)PbCl4 has the broadest emission band centered at 520

    nm, with FWHM of ~160 nm (Fig. 2c). Based on emission wavelength, MAPbBr3 and

    (EDBE)PbCl4 appear to be the most promising candidates for the scintillators coupled to APD

    [28].

    In all perovskite crystals, X-ray excited luminescence and photoluminescence spectra are

    very similar, indicating that the dominant scintillation mechanism is straightforward: upon X-ray

    absorption, high-energy excitations thermalize through ionizations and excitations of atoms,

    until excitons are generated at energies near the bandgap. X-ray excited luminescence stems

    solely from the intrinsic excitonic emission of the perovskites, and no other defect states seem to

    be involved in the scintillation process.

    The dynamics of radiative processes in materials under high-energy excitation is often

    complicated by slower non-exponential components due to charge carrier trapping and re-

    trapping, which manifest themselves as delayed luminescence, or afterglow. Upon termination of

    the X-ray excitation, afterglow effects would typically contribute a residual luminescence

    background with characteristic lifetime of few ms, thus lowering the effective light yield and

    worsening the signal-to-noise ratio. Afterglow effects are particularly detrimental for

    applications like computed tomography, in which temporal crosstalk considerably reduces the

    image quality [2]. Charge carrier trapping and re-trapping processes can be monitored by

    thermoluminescence measurements. In our specific mode of operation for thermoluminescence

    intensity measurements, we were able to record steady-state X-ray excited luminescence

  • 9

    intensity during irradiation, immediately prior to the thermoluminescence scan (see details in

    Materials and methods). In this way, two distinct integrated intensities can be evaluated: the first

    one, which we denote as ITL, comprising the range from the end of X-ray irradiation till the end

    of the entire run, while the second one, denoted as ITL+IssXL, comprising the range from the start

    of the X-ray irradiation until the end of the run. This allows calculating, for each sample, the

    ITL/(ITL+IssXL) ratio, which can be interpreted as the fraction of the total excitation energy

    accumulated into traps [19,32]. The value of this ratio, therefore, provides a qualitative estimate

    of the influence of traps on the scintillation yield.

    Figure 3. Residual luminescence background after X-ray excitation. Low

    temperature thermoluminescence curves of a) MAPbI3, b) MAPbBr3, and c)

    (EDBE)PbCl4. The data are presented on a time scale starting at temperature of 10 K and

    increasing to 350 K after 3600 s, as indicated by the dashed line in the right panel

    (temperature scale on the right axes).

    Typical thermoluminescence curves of the metal halide perovskite crystals are shown as solid

    curves in Fig. 3. After termination of the X-ray excitation at 10 K, long tails extending to

    thousands of seconds were observed in all crystals. Although the long-lived component of this

  • 10

    afterglow effect is much slower than the photoluminescence decay (see Supplementary Fig. S6),

    it only occurs at low temperatures (~10 K) and is negligible at room temperature. In the case of

    MAPbI3 and MAPbBr3, low temperature thermoluminescence curves are dominated by a double-

    structured peak, with two smaller satellite peaks appearing at longer times (Figs. 3a and 3b). In

    (EDBE)PbCl4, the low-temperature thermoluminescence curve shows that one peak strongly

    dominates the other peak while the total intensity of the peaks is much higher than those in

    MAPbI3 and MAPbBr3 (Fig. 3c). The ratio of ITL/(ITL+IssXL) ~ 0.002 is very similar in both

    MAPbI3 and MAPbBr3, which is extremely low in comparison with other oxide materials used

    for scintillators, such as lanthanide aluminium perovskite or garnets [19,32,33,35]. Moreover,

    MAPbI3 and MAPbBr3 crystals show nearly trap-free behavior from T=75 K up to the highest

    temperature investigated of T=350 K, a very desirable characteristic from the point of view of

    scintillation speed and efficiency. In (EDBE)PbCl4, ITL/(ITL+IssXL) ~ 0.058, a slightly higher

    value than in the three-dimensional perovskite crystals, but still relatively low.

    Figure 4. Determination of low-energy trap states. Glow curves of a) MAPbI3, b)

    MAPbBr3, and c) (EDBE)PbCl4 recorded after 10 min X-ray irradiation at 10 K, at a

    heating rate of 0.14 K/s. The solid lines are the best fits to the experimental data points by

    multiple Randall-Wilkins equations (Eq. 1): single components and peak temperatures

    (Tmax) are indicated by dashed lines and arrows, respectively (see Table 1 for fitting

    parameters).

  • 11

    The zero-order glow curves of the three crystals are presented in Fig. 4. Appearance of

    thermoluminescence signal at temperatures below 150 K reveals the existence of low-energy trap

    states. Since for such states it is difficult to determine the exact number of traps, their depth and

    frequency factors [33], we restrict our analysis to thermoluminescence peaks with intensity

    larger than 10 % of the maximum. Thermoluminescence curves have been deconvoluted into k

    glow peaks, based on the classic Randall-Wilkins equation [34]:

    𝐼𝑇𝐿 = ∑ 𝑛0𝑖𝑉𝜎𝑖exp (−𝐸𝑖

    𝑘𝐵𝑇)

    𝑘

    𝑖=1

    exp (−𝜎𝑖𝛽

    ∫ exp (−𝐸𝑖

    𝑘𝐵𝑇′) 𝑑𝑇′

    𝑇

    𝑇0

    ) (1)

    where T is the temperature, β the heating rate, and kB the Boltzmann constant; n0i is the initial

    trap concentration, V is the crystal volume, Ei the trap depth, σi the frequency factor of each

    component. Note that the unitless number of traps n0iV is often used to compare the afterglow of

    different crystals [19,32,33,34,35].

    This analysis provides a good indication of the characteristics of prevailing trap states,

    however it cannot resolve the existence of traps that fall at times much longer than seconds, or

    with mixed order kinetics [32]. The room temperature lifetime of trapped carriers, such as

    electron or hole centers and excitons, τi, can also be estimated from the energy and frequency

    factor of the trap, using the well-known Arrhenius formula:

    1

    𝜏𝑖= 𝜎𝑖exp (−

    𝐸𝑖𝑘𝐵𝑇

    ) (2)

    While the glow curves of MAPbI3 and MAPbBr3 in Fig. 4a and b have been fitted using four

    and three components, respectively, the glow curve of (EDBE)PbCL4 in Fig. 4c could be fitted

    with only two components. The corresponding fitting parameters are shown in Table 1. All

  • 12

    crystals have relatively low trap densities, with depth energy (E) varying from ~10 to 90 meV.

    The initial trap concentrations n0 in MAPbI3 and MAPbBr3 can be calculated from the total

    number of traps (n0V~103-104) and the volume of the crystal (V~30-100 mm3). The resulting trap

    concentrations (n0~105-107 cm-3) are comparable to those of shallow traps previously observed in

    photoconductivity measurements (~105-107 cm-3) [11] and space-charge-limited-current (~109-

    1010 cm-3) [23], also considering the uncertainty in the estimate of the active crystal volume. The

    fastest room temperature lifetimes (τ) of MAPbI3 and MAPbBr3 are of the order of milliseconds,

    long enough to contribute to the light yield components without residual luminescence

    background. Correspondingly, logarithmic frequency factors (ln σ) are all below 16, which is

    much smaller than ln σ~30 typically found in pristine or activated oxide materials [19,32,33,35],

    also reported in Table 1 for comparison. (EDBE)PbCl4 has the largest trap density among the

    perovskites we investigated, n0~107 cm-3. Large concentration of shallow traps may be beneficial

    for X-ray scintillation at low-temperature, as in the case of Ce3+-doped YAlO3 and LuAlO3 [35],

    or pristine Li2B4O7 [36]. This is indeed seen in temperature dependent X-ray excited

    luminescence spectral maps shown in Fig. 5.

  • 13

    Compound

    Tmax (K) E (eV) ln σ (s-1) τ (s) n0V Reference

    MAPbI3

    32 0.0309 8.09 1.04⋅10-3 2.45⋅104

    This work 46 0.0226 1.78 0.41 1.85⋅104

    56 0.0901 15.60 5.95⋅10-4 6.12⋅103

    62 0.0389 3.25 0.18 1.45⋅104

    MAPbBr3

    33 0.0139 1.16 0.54 7.61⋅104 This work 56 0.0602 9.02 1.31⋅10-3 2.10⋅104

    68 0.0909 12.1 2.04⋅10-4 2.73⋅104

    EDBEPbCl4 32 0.0177 2.83 0.12 5.95⋅105 This work

    45 0.0281 3.40 0.10 1.71⋅106

    LuAlO3: Ce3+

    36 0.0148 0.9346 2.16⋅10-2 2.84⋅104

    [19,32,35]

    88 0.105 10.07 2.29⋅10-2 1.53⋅104

    187 0.498 27.22 2.51⋅10-2 2.10⋅106

    206 0.385 17.56 1.61⋅10-2 4.64⋅104

    223 0.669 30.99 2.18⋅10-2 1.38⋅104

    253 0.75 30.53 2.08⋅10-2 4.84⋅104

    273 0.799 30.08 2.05⋅10-2 1.52⋅105

    YAlO3: Ce3+

    108 0.30 29.24 4.99⋅10-2 ~105 [19,32,35] 154 0.50 34.18 3.02⋅10-2 ~105

    281 0.421 18.05 1.95 ~104

    Gd3Al2Ga3O12: Ce3+

    36 0.0576 15.9 1.2⋅10-6 1.6⋅105

    [33]

    45 0.0446 8.32 1.4⋅10-3 4.7⋅105

    73 0.116 15.1 2.7⋅10-5 3.4⋅105

    181 0.211 9.01 0.52 2.9⋅105

    240 0.527 21.31 0.65 2⋅105

    255 0.321 9.76 19 7.5⋅105

    Table 1. Trap state parameters. The parameters were derived from the fitting of first-

    order glow curves in Figure 4: Tmax is the temperature at which the glow curve peaks, E

    the trap depth, ln σ the logarithmic frequency factor in s-1, τ the room temperature

    lifetime, and n0V the total, initial number of traps. Comparative parameters of known

    scintillator materials from the literature are also reported in the last three lines.

  • 14

    Figure 5. Temperature dependent luminescence under X-ray excitation. X-ray

    excited luminescence spectra (X-ray excited luminescence) of perovskite crystals at

    various temperatures, from 10 to 350 K: (a) MAPbI3, (b) MAPbBr3, and (c)

    (EDBE)PbCl4. (d) Comparison of the normalized X-ray excited luminescence spectra at

    T=10 K.

    MAPbI3 (Fig. 5a) and MAPbBr3 (Fig. 5b) show strong dependence of X-ray excited

    luminescence on temperature, with significantly reduced emission at temperatures above 100 K.

    At very low temperatures they display distinct emission bands with sharp maxima at 770 nm and

    540 nm, respectively (see Fig. 5d for comparison of X-ray excited luminescence spectra recorded

  • 15

    at T=10 K). Notably, the X-ray excited luminescence peak at 770 nm, with FWHM of 5 nm, has

    the same characteristics of coherent light emission previously observed in MAPbI3 [17]. Side

    bands also appear at 760 and 800 nm in MAPbI3 and at ~590 nm in MAPbBr3, but their origin is

    still unclear. The X-ray excited luminescence spectrum of (EDBE)PbCl4 (Fig. 5c) consists of a

    much broader band peaking at ~520 nm, with intensity significantly less sensitive to temperature.

    As temperature increases, the X-ray excited luminescence intensity first decreases between 10

    and 50 K, then increases towards 130 K, and reduces steadily at higher temperatures. In all

    crystals, the FWHM of X-ray excited luminescence peaks increases with increasing temperature,

    consistent with the spreading of excited electrons to high vibrational levels [37].

    As discussed previously, the light yield of perovskite single crystals estimated from their

    bandgaps should be >120,000 photons/MeV. From the pulse height spectra of samples excited

    with 662 keV γ-ray of Cs137 shown in Supplementary Fig. S7, the actual light yield of

    (EDBE)PbCl4 at room temperature is moderate, with ~9,000 photons/MeV. We note that they

    are not so many reports about the energy spectra from γ-ray reported for perovskite scintillator

    [10] and direct conversion detector [12]. Light yield of (EDBE)PbCl4 is actually similar to that

    of two-dimensional perovskite crystal PhE-PbBr4 (10,000 photons/MeV) reported previously

    [31]. The light yields of MAPbBr3 and MAPbI3 at room temperature are much lower, and cannot

    be extracted from pulse height experiments. Low light yields at room temperature may arise

    from thermally activated quenching effects. To confirm this hypothesis, we have derived light

    yields at different temperatures from the integrated intensities of the X-ray excited emission

    spectra in Fig. 5; considering the small afterglow, we expect the light yield derived from X-ray

    excited emission spectra to be very similar to that derived from pulse height spectra.

  • 16

    Figure 6. Temperature dependence of the light yields. Light yields of MAPbI3,

    MAPbBr3, and (EDBE)PbCl4 obtained from the integrated X-ray excited luminescence

    intensities at various temperatures, from 10-350 K. The left axis shows integrated

    intensity in arbitrary units obtained from the corrected X-ray excited luminescence

    spectra in Fig. 5, while the right axis exhibits the light yield in absolute units after

    calibration of the light yield of (EDBE)PbCl4 at 300 K to ~9,000 photons/MeV, as

    derived independently from its pulse height spectrum. The inset shows details of the

    curves from 290 to 310 K.

    Light yields derived from the integrated X-ray excited luminescence emission intensities of

    the three halide perovskite crystals as a function of temperature are reported in Fig. 6. We

    integrated the corrected intensity of X-ray excited luminescence spectra in Fig. 5 and used the

    light yield of ~9,000 photons/MeV derived from the pulse height spectrum of (EDBE)PbCl4 at

    300 K (Supplementary Fig. S7) to calibrate the integrated intensity. For (EDBE)PbCl4, the

    resulting light yield at 300 K is about ~8% of the maximum at 130 K. Since the light yield is

    linearly proportional to the photoluminescence quantum efficiency [18] while the efficiency of

    charge transport to the recombination center is almost unity [24,29], the ratio of 8% is consistent

  • 17

    with reported (EDBE)PbCl4 photoluminescence quantum efficiency of less than 10% at room

    temperature. Light yields of MAPbI3 and MAPbBr3 are

  • 18

    materials [19,33]. Low-temperature X-ray excited luminescence measurements have shown that

    the X-ray luminescence yield can be as high as ~120,000 photons/MeV in (EDBE)PbCl4 at

    T=130 K, and in excess of 150,000 photons/MeV in MAPbI3 and MAPbBr3 at T=10 K. The wide

    synthetic versatility of hybrid perovskites allows easy tuning of their scintillation properties: for

    example, their emission spectra can be controlled by cation or halide substitution to perfectly

    match the spectral sensitivity of high-quantum-efficiency APD, like in the case of MAPbBr3 and

    (EDBE)PbCl4. Their emissive properties can be further enhanced through engineering of

    perovskite structure and dimensionality: while light yield of 3D perovskites MAPbI3 and

    MAPbBr3 is significantly reduced at room temperature (

  • 19

    and 350 K. We note that the measurements were carried out from 350 K to 10 K to avoid

    possible contributions from thermal release of charge carriers to the emission yield. After X-ray

    excited luminescence measurements, we measured low temperature thermoluminescence and

    glow curves. Prior to thermoluminescence measurements, the samples were exposed to X-rays

    for 10 min at 10 K. Thermoluminescence and glow curves were recorded between 10 and 350 K

    at a heating rate of about 0.14 K/s. Thermoluminescence curves were recorded with the

    monochromator set to the zeroth order. Photoluminescence spectra were recorded with a

    commercial spectrofluorometer HORIBA Jobin Yvon Fluorolog-3 spectrofluorometer at room

    temperature.

    Crystal growth: Three-dimensional perovskite precursors, MABr and MAI, were synthetized by

    mixing hydrobromic acid (48% wt in water, Sigma-Aldrich) and hydroiodic acid (57% wt in

    water, Sigma-Aldrich) with methylamine solution (CH3NH2 , 40% in methanol, Tokyo Chemical

    Industry, Co., Ltd) in 1:1 molar ratio. The ice-cooled mixture was left under magnetic stirring for

    2 h, and the solvent removed with a rotary evaporator. The resulting powders were dissolved in

    ethanol, crystallized with diethylether for purification repeating the cycle 6 times, and finally

    dried in vacuum oven at 6 ○C for 12 h. For (EDBE)PbCl4(EDBE=2,2’-

    (ethylenedioxy)bis(ethylammonium)), the organic precursor (EDBE)Cl2 was synthetized in

    aqueous solution by reaction of 2,2’-(ethylenedioxy)bis(ethylamine) (98%, Sigma Aldrich) with

    excess of HCl (37% in H2O). The solution was stirred for 4h at room temperature to complete

    the reaction. A purification process similar to that discussed for MABr and MAI was applied to

    collect the final white and high purity powders.

    For the synthesis of hybrid perovskite crystals, the following inorganic precursors were

    purchased from Sigma-Aldrich: lead(II) chloride (PbCl2, 99.999%), lead(II) bromide (PbBr2,

  • 20

    99.999%) and lead(II) iodide (PbI2, 99.0%). Crystals of MAPbBr3 were synthetized using inverse

    temperature crystallization as similarly reported elsewhere [27]. 2 ml of 1M DMF solution of

    MABr and PbBr2 (1:1 molar ratio) were left overnight on a hotplate (110 ○C) without stirring,

    allowing the precipitation of the perovskite crystals. MAPbI3 were obtained by slow evaporation

    at room temperature of a saturate N, N-dimethylformamide (DMF) solution of MAI and PbI2

    (1:1 molar ratio). To obtain (EDBE)PbCl4 crystals, a 1M solution of (EDBE)Cl2 and PbCl2 (1:1

    molar ratio) in dimethylsulphoxide (DMSO) was prepared by dissolving the precursors at 110 ○C

    on a hotplate. After natural cooling of the solution at room temperature, slow crystallization over

    a period of 1 month results in the formation of cm-scale white perovskite crystals. The

    crystallization processes were performed under inert N2 atmosphere. All the obtained crystals

    were collected from the precursor solutions, washed with diethylether and dried in vacuum

    overnight.

    Acknowledgements

    Research was supported by the Ministry of Education (MOE2013-T2-044 and MOE2011-T3-1-

    005) and by the National Research Foundation (NRF-CRP14-2014-03) of Singapore. X-ray

    excited luminescence and thermoluminescence measurements were performed at the National

    Laboratory for Quantum Technologies (NLTK), Nicolaus Copernicus University, supported by

    the European Regional Development Fund.

    Author contributions

    MDB and CS conceived the idea. DC synthesized the perovskite precursors, prepared, and

    characterized crystals and films. MDB and WD designed the experiments. KB and WL

    performed X-ray excited luminescence and thermoluminescence measurements. DC and AB

  • 21

    collected absorption and photoluminescence measurements of thin films. WD, MDB, DC and CS

    analyzed the data. MDB and CS drafted the manuscript. All the authors contributed to

    interpretation of the results and revision of the manuscript. CS supervised the work. All authors

    take full responsibility for the content of the paper.

    Additional information

    Supplementary information provided: powder X-ray diffraction measurements of the perovskite

    crystals to confirm the expected perovskite structures; absorption and photoluminescence

    measurements of perovskite thin films as a reference for the optical properties of the crystals;

    calculation of X-ray absorption lengths of the three perovskite crystals, which turn out to be

    comparable to those of commercial LaBr3 and LuI3 scintillators; time-resolved

    photoluminescence of single crystals as some decay components are much faster than those in

    commercial scintillators; pulse height spectra that provide additional details on scintillation

    properties.

    The authors declare that they have no competing financial interests. Reprints and permission

    information is available online at http://npg.nature.com/reprintsandpermissions/. Correspondence

    and requests for materials should be addressed to MDB ([email protected]) and CS

    ([email protected]).

    References

    [1] Blasse, G., Scintillator materials, Chem. Mater. 6, 1465-1475 (1994).

    [2] Nikl, M., Scintillation detectors for x-rays, Meas. Sci. Technol. 17, R37-R54 (2006).

    [3] Street, R. A., Ready, S. E., Van Schuylenbergh, K., Ho, J., Boyce, J. B., Nylen, P., Shah, K.,

    Melekhov, L., and Hermon, H., Comparison of PbI2 and HgI2 for direct detection active matrix

    x-ray image sensors, J. Appl. Phys. 91, 3345-3355 (2002).

  • 22

    [4] Szeles, C., CdZnTe and CdTe materials for X-ray and gamma ray radiation detector

    applications, phys. stat. sol. (b) 241, 783-790 (2004).

    [5] Kasap, S. and Rowlands, J., Direct-conversion flat-panel X-ray image sensors for digital

    radiography, Proc. IEEE 90, 591-604 (2002).

    [6] Büchele, P., Richter, M., Tedde, S. F., Matt, G. J., Ankah, G. N., Fischer, R., Biele, M.,

    Metzger, W., Lilliu, S., Bikondoa, O., Macdonald, J. E., Brabec, C. J., Kraus, T., Lemmer, U.,

    and Schmidt, O., X-ray imaging with scintillator-sensitized hybrid organic photodetectors,

    Nature Photon. 9, 843-848 (2015).

    [7] Heiss, W. and Brabec, C., X-ray imaging: Perovskites target X-ray detection, Nature Photon.

    10, 288-289 (2016).

    [8] Tegze, M. and Faigel, G., X-ray holography with atomic resolution, Nature 380, 49-51

    (1996).

    [9] Rieder, R., Economou, T., Wanke, H., Turkevich, A., Crisp, J., Bruckner, J., Dreibus, G., and

    McSween, H., The chemical composition of Martian soil and rocks returned by the mobile alpha

    proton x-ray spectrometer: Preliminary results from the x-ray mode, Science 278, 1771-1774

    (1997).

    [10] Yakunin, S., Sytnyk, M., Kriegner, D., Shrestha, S., Richter, M., Matt, G. J., Azimi,

    H.,Brabec, C. J., Stangl, J., Kovalenko, M. V., and Heiss, W., Detection of X-ray photons by

    solution-processed lead halide perovskites, Nature Photon. 9, 444-450 (2015).

    [11] Wei, H., Fang, Y., Mulligan, P., Chuirazzi, W., Fang, H.-H., Wang, C. , Ecker, B. R., Gao,

    Y., Loi, M. A., Cao, L., and Huang, J., Sensitive X-ray detectors made of methylammonium lead

    tribromide perovskite single crystals, Nature Photon. 10, 333-339 (2016).

    [12] Yakunin, S., Dirin, D. N., Shynkarenko, Y., Morad, V., Cherniukh, I., Nazarenko, O.,

    Kreil, D., Nauser, T., and Kovalenko, M. V., Detection of gamma photons using solution-grown

    single crystals of hybrid lead halide perovskites, Nature Photon. AOP (2016).

    [13] Chin, X.Y., Cortecchia, D., Yin, J., Bruno, A., and Soci, C., Lead iodide perovskite light-

    emitting field-effect transistor, Nature Comm. 6, 7383-1-7383-9 (2015).

    [14] Dou, L., Yang, Y. M., You, J., Hong, Z., Chang, W.H., Li, G., and Yang, Y., Solution-

    processed hybrid perovskite photodetectors with high detectivity, Nature Comm. 5, 5404-1-

    5404-6 (2014).

  • 23

    [15] Chondroudis, K. and Mitzi, D. B., Electroluminescence from an organic−inorganic

    perovskite incorporating a quaterthiophene dye within lead halide perovskite layers, Chem.

    Mater. 11, 3028-3030 (1999).

    [16] Tan, Z.-K., Moghaddam, R. S., Lai, M. L., Docampo, P., Higler, R., Deschler, F., Price, M.,

    Sadhanala, A., Pazos, L. M., Credgington, D., Hanusch, F., Bein, T., Snaith, H. J., and Friend, R.

    H., Bright light-emitting diodes based on organometal halide perovskite, Nature Nanotech. 9,

    687-692 (2014).

    [17] Xing, G., Mathews, N., Lim, S. S., Yantara, N., Liu, X., Sabba, D., Grätzel, M., Mhaisalkar,

    S.and Sum, T. C., Low-temperature solution-processed wavelength-tunable perovskites for

    lasing, Nature Mater. 13, 476-480 (2014).

    [18] Birowosuto, M. D. and Dorenbos, P., Novel γ- and X-ray scintillator research: on the

    emission wavelength, light yield and time response of Ce3+ doped halide scintillators, phys. stat.

    sol. (a) 206, 9-20 (2009).

    [19] Drozdowski, W., Wojtowicz, A. J., Lukasiewicz, T., and Kisielewski, J., Scintillation

    properties of LuAP and LuYAP crystals activated with Cerium and Molybdenum, Nucl. Instr.

    Meth. Phys. Res. A 562, 254-261 (2006).

    [20] Shibuya, K., Koshimizu, M., Takeoka, Y., and Asai, K., Scintillation properties of

    (C6H13NH3)(2)PbI4: Exciton luminescence of an organic/inorganic multiple quantum well

    structure compound induced by 2.0 MeV protons, Nucl. Instr. Meth. Phys. Res. B 194, 207-212

    (2002).

    [21] Kishimoto, S., Shibuya, K., Nishikido, F., Koshimizu, M., Haruki, R., and Yoda, Y.,

    Subnanosecond time-resolved x-ray measurements using an organic-inorganic perovskite

    scintillator, Appl. Phys. Lett. 93, 261901-1-261901-3 (2008).

    [22] Saidaminov, M. I., Abdelhady, A. L., Murali, B., Alarousu, E., Burlakov, V. M., Peng, W.,

    Dursun, I., Wang, L., He, Y., Maculan, G., Goriely, A., Wu, T., Mohammed, O. F., and Bakr, O.

    M., High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature

    crystallization, Nature Comm. 6, 7586-1-7586-6 (2015).

    [23] Shi, D., Adinolfi, V., Comin, R., Yuan, M., Alarousu, E., Buin, A., Chen, Y., Hoogland, S.,

    Rothenberger, A., Katsiev, K., Losovyj, Y., Zhang, X., Dowben, P. A., Mohammed, O. F.,

    Sargent, E. H., and Bakr, O. M., Low trap-state density and long carrier diffusion in organolead

    trihalide perovskite single crystals, Science 347, 519-522 (2015).

  • 24

    [24] Dohner, E. R., Jaffe, A., Bradshaw, L. R., and Karunadasa, H. I., Intrinsic white-light

    emission from layered hybrid perovskites, J. Am. Chem. Soc. 136, 13154-13157 (2014).

    [25] van Loef, E. V. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. and Güdel, H. U., High-

    energy-resolution scintillator: Ce3+ activated LaBr3, Appl. Phys. Lett. 79, 1573-1-1573-3 (2001).

    [26] Birowosuto, M. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. W., and Güdel, H. U.,

    Thermal quenching of Ce3+ emission in PrX3 ( X = Cl , Br ) by intervalence charge transfer, J.

    Phys. Condens. Matter 19, 256209-1-256209-16 (2007).

    [27] Birowosuto, M. D., Dorenbos, P.. de Haas, J. T. M., van Eijk, C. W. E., Krämer, K. W., and

    Güdel, H. U., Optical spectroscopy and luminescence quenching of LuI3: Ce3+, J. Lumin. 118,

    308-316 (2006).

    [28] Birowosuto, M. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. W., and Güdel, H. U.,

    High-light-output scintillator for photodiode readout: LuI3: Ce3+, J. Appl. Phys. 99, 123520-1-

    123520-4 (2006).

    [29] Cortecchia, D., Yin, J., Lova, P., Mhaisalkar, S., Gurzadyan, G. G., Bruno, A., and Soci, C.,

    Polaron self-localization in white-light emitting hybrid perovskites, arXiv:1603.01284 [cond-

    mat.mtrl-sci] (2016).

    [30] Birowosuto, M. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. W., and Güdel, H. U.,

    PrBr3: Ce3+: A New Fast Lanthanide Trihalide Scintillator, IEEE Trans. Nucl. Sci. 53, 3028-3030

    (2006).

    [31] van Eijk, C. W. E., de Haas, J. T. M., Rodnyi, P. A., Khodyuk, I. V., Shibuya, K.,

    Nishikido, F., and Koshimizu, M., Scintillation properties of a crystal of

    (C6H5(CH2)2NH3)2PbBr4, IEEE Nuclear Science Symposium Conference Record 2008, 3525-

    3528 (2008).

    [32] Bartram, R. H., Hamilton, D. S., Kappers, L., and Lempicki, A., Electron traps and transfer

    efficiency of cerium-doped aluminate scintillators, J. Lumin. 75, 183-192 (1997).

    [33] Drozdowski, W., Brylew, K., Witkowski, M. E., Wojtowicz, A. J., Solarz, P., Kamada, K.,

    and Yoshikawa, A., “Studies of light yield as a function of temperature and low temperature

    thermoluminescence of Gd3Al2Ga3O12: Ce scintillator crystals, Opt. Mater. 36, 1665-1669

    (2014).

    [34] Randall, J., and Wilkins, M., The phosphorescence of various solids, Proc. Roy. Soc.

    London A 184, 366-407 (1945).

  • 25

    [35] Wojtowicz, A. J., Glodo, J., Drozdowski, W., and Przegietka, K. R., Electron traps and

    scintillation mechanism in YAlO3: Ce and LuAlO3: Ce scintillators J. Lumin. 79, 275-291

    (1998).

    [36] Ogorodnikov, I. N. and Poryvai, N. E., Thermoluminescence kinetics of lithium borate

    crystals, J. Lumin. 132, 1318-1324 (2012).

    [37] Liu, C., Qi, Z., Ma, C.-g., Dorenbos, P., Hou, D., Zhang, S., Kuang, X., Zhang, J., and

    Liang, H., High light yield of Sr8(Si4O12)Cl8: Eu2+ under x-ray excitation and its temperature-

    dependent luminescence characteristics Chem. Mater. 26, 3709–3715 (2014).

    [38] Sutter-Fella, C. M., Li, Y., Amani, M., Ager III, J. W., Toma, F. M., Yablonovitch, E.,

    Sharp, I. D., and Javey, A., High photoluminescence quantum yield in band gap tunable bromide

    containing mixed halide perovskites, Nano Lett. 16, 800-806 (2016).

    [39] Kolk, E. V. D. and Dorenbos, P., Systematic and material independent variation of

    electrical, optical ,and chemical properties of Ln materials over the Ln series ( Ln ) La, Ce, Pr, ..,

    Lu, Chem. Mater. 18, 3458-3462 (2006).

    [40] Birowosuto, M. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. W., and Güdel, H. U.,

    Ce3+ activated LaBr3-xIx : High-light-yield and fast-response mixed halide scintillators, J. Appl.

    Phys. 103, 103517-1-103517-6 (2008).

  • 26

    Supplementary Information

  • 27

    X-ray Diffraction

    To characterize the crystal structure of the perovskite crystals, X-ray powder diffraction (XRPD)

    was performed on perovskite powders obtained from ground crystals on a BRUKER D8

    ADVANCE with Bragg-Brentano geometry using Cu Kα radiation (λ = 1.54,056 Å), step increment of 0.02○ and 1 s of acquisition time. The results are shown in Figs. S1, S2 and S3.

    Figure S1. XRD pattern of MAPbBr3 powders. The diffractogram is consistent

    with the perovskite structure having cubic crystal system, space group Pm3m and

    lattice parameters a = 5.917(1) Å.

    Figure S2. XRD pattern of MAPbI3 powders. The diffractogram is consistent

    with the perovskite structure having tetragonal crystal system, space group I4 ⁄

    mcm and lattice constants a = 8.867(5) Å and b = 12.649(3) Å.

  • 28

    Figure S3. XRD pattern of (EDBE)PbCl4 powders. The 00l reflections indicate

    the formation of the layered structure of the two-dimensional perovskite, in

    agreement with the previously reported structure (by Dohner et al, [1] having

    monoclinic crystal system, space group C2 and refined lattice parameters a =

    7.762(8) Å, b = 7.629(2) Å, c = 13.375(7) Å, β = 102.7(2) Å.

    Absorption and Photoluminescence

    Absorption and photoluminescence measurements were performed in order to obtain the energy

    band gap and confirm the large Stoke shift in two-dimensional perovskite scintillators (Fig. S4).

    Absorption spectra were recorded by an UV-VIS-NIR spectrophotometer (UV3600, Shimadzu)

    using a scanning resolution of 0.5 nm. Steady-state photoluminescence spectra were recorded by

    a Fluorolog-3, (HORIBA Jobin Yvon) spectrofluorometer with wavelength resolution 0.5 nm.

    Figure S4. Room temperature absorption (blue) and photoluminescence spectrum (red)

    of a) MAPbI3, b) MAPbBr3, c) (EDBE)PbCl4 (thin films). The photoluminescence peak

    blue-shifts from 760 nm to 527 nm from MAPbI3 to MAPbBr3, following the

    corresponding blue-shift of the absorption edges. The absorption spectrum of

    (EDBE)PbCl4 shows a pronounced excitonic peak at 326 nm and broadband, highly

    Stoke-shifted photoluminescence peaked at 525 nm.

  • 29

    Absorption Length

    Three types of the interaction mechanisms for electromagnetic radiation in matter play an

    important role in the absorption of X- and -rays. These are photoelectric absorption, Compton

    scattering, and pair production. All these processes lead to the partial or complete absorption of

    the radiation quantum. The absorption length was obtained from formula [2]:

    𝑙𝑎𝑏𝑠 ≈ 2𝑍𝑒𝑓𝑓

    𝑁𝐴

    1

    𝜌𝜎

    where Zeff, NA, , and are the effective atomic number, Avogadro number, the mass density, and the absorption cross section for each atomic element. As the cross section was separately

    characterized for photoelectric, Compton scattering, and pair production, the total absorption

    length was determined by the inverse sum of the absorption lengths for the three interaction

    mechanisms mentioned above (Fig. S5).

    Figure S5. Calculated absorption length of perovskite crystals as a function of

    photon energy, covering the X-ray spectral region. Curves for Ce3+ doped LuI3 [2]

    and LaBr3 [3] scintillators are also added for comparisons.

    Time Resolved Photoluminescence

    The microphotoluminescence setup is based on free space excitation technique with the

    excitation path and the emission collection from the side, using a VIS-NIR microscope objective

    (40x, NA=0.65). The MAPbI3, MAPbBr3, and (EDBE)PbCl4 single crystals were excited with 5-

    MHz-repetition-rate, picosecond-pulse light sources at 640 and 370 nm of Edinburgh laser

    diodes and at 330 nm of a Picoquant light-emitting diode, respectively. In all cases, the beam

    spot size was about 2 mm. A silicon-based charge-coupled-device camera was used for imaging.

    Time-resolved decay curves were obtained using grating Edinburgh Instruments or tunable

    bandpass filters at 766, 540, and 520 nm for MAPbI3, MAPbBr3 and (EDBE)PbCl4 crystals,

    respectively. The signal from the Hamamatsu photomultiplier or Micro Photon Devices single-

    photon avalance photodiode was acquired by a time-correlated single photon counting card.

  • 30

    Figure S6. Time resolved photoluminescence curves of MAPbI3, MAPbBr3 and

    (EDBE)PbCl4 single crystals. Excitation and emission wavelengths are reported in the

    text. All intensities were normalized, and the one of (EDBE)PbCl4 was further divided by

    a factor of five for clarity. The white lines in the curves are exponential fittings of the

    data.

    The decay curves of MAPbI3, MAPbBr3 and (EDBE)PbCl4 were fitted with double, triple, and

    single exponential fits, respectively. The resulting decay components of MAPbI3 are 4.3 and 52.2

    ns, with contributions of 18 and 82 %, respectively. Those of MAPbBr3 are 0.8, 5.2, and 45.4 ns

    with contributions of 10, 18 and 72 %, respectively. While the longer decay times are consistent

    with the values previously reported for these 3D perovskites [4], our instrumental resolution

    (0.05 ns) allowed to resolve the additional presence of the fast components with decay times < 1

    ns. Due to the limited time window, ultralong-lived components (> 300 ns) were not detected.

    Finally, (EDBE)PbCl4 decay has only one component of 7.9 ns, consistent with the

    photoluminescence lifetime reported for similar 2D perovskites [1]. Note that all the fast

    components are below 10 ns, much faster than those of commercial scintillators based on Ce3+

    doped LuI3 [2] and LaBr3 [3].

    Pulse Height Spectra

    Pulse height spectra were measured at room temperature under 662 keV gamma excitation from

    a 137Cs source (no. 30/2010, 210 kBq). The pulsed output signal from a Hamamatsu R2059

    photomultiplier was processed by a Canberra 2005 integrating preamplifier, a Canberra 2022

    spectroscopy amplifier, and a multichannel analyzer. To improve the light collection efficiency

    the samples were coupled to the quartz window of the photomultiplier with Viscasil grease and

    covered with several layers of Teflon tape. Light yield is obtained from the position of the 662

    keV photopeak in pulse height spectra both recorded with the photomultiplier and the APD.

    Using the photomultiplier, the photoeletron yield, expressed in photoelectrons per MeV of

    absorbed -ray energy (phe/MeV), is determined by comparing of the peak position of the 662

    keV photopeak to the position of the mean value of the single electron response [2].

  • 31

    Figure S7. Pulse height spectra of perovskite crystals under 662 keV of 137Cs

    source with a pulse shaping time of 2 s. The room temperature light yields

    derived for (EDBE)PbCl4 and MAPbBr3 are 9,000 and < 1,000 photons/MeV,

    respectively.

    References

    [1] Dohner, E. R., Jaffe, A., Bradshaw, L. R., and Karunadasa, H. I., Intrinsic white-light emission from layered hybrid perovskites, J. Am. Chem. Soc. 136, 13154-13157 (2014).

    [2] Birowosuto, M. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. W., and Güdel, H. U.,

    High-light-output scintillator for photodiode readout: LuI3: Ce3+, J. Appl. Phys. 99,

    123520-1-123520-4 (2006).

    [3] van Loef, E. V. D., Dorenbos, P., van Eijk, C. W. E., Krämer, K. and Güdel, H. U.,

    High-energy-resolution scintillator: Ce3+ activated LaBr3, Appl. Phys. Lett. 79, 1573-1-

    1573-3 (2001).

    [4] Shi, D., Adinolfi, V., Comin, R., Yuan, M., Alarousu, E., Buin, A., Chen, Y., Hoogland,

    S., Rothenberger, A., Katsiev, K., Losovyj, Y., Zhang, X., Dowben, P. A., Mohammed,

    O. F., Sargent, E. H., Bakr, O. M., Low trap-state density and long carrier diffusion in

    organolead trihalide perovskite single crystals, Science 347, 519 (2015).